검색
Article Search

JMB Journal of Microbiolog and Biotechnology

OPEN ACCESS eISSN 0454-8124
pISSN 1225-6951
QR Code

Article

Kyungpook Mathematical Journal 2024; 64(4): 559-578

Published online December 31, 2024 https://doi.org/10.5666/KMJ.2024.64.4.559

Copyright © Kyungpook Mathematical Journal.

Synchronization Estimate of the Discrete Kuramoto Model with Multiplicative Random Noise

Woojoo Shim

Department of Mathematics Education, Kyungpook National University, Daegu 41566, Republic of Korea
e-mail: wjshim@knu.ac.kr

Received: September 7, 2024; Accepted: September 13, 2024

We study the onset of synchronization in the time-discretization of the Kuramoto model with multiplicative noise. Specifically, we present a sufficient condition that ensures that, starting from generic initial data, identical oscillators in the discrete-time Kuramoto model with multiplicative noise will be asymptotically synchronized almost surely.

Keywords: Kuramoto model, order parameter, synchronization, Euler- Maruyama method

Synchronization is a unique form of emergent behavior frequently observed in nature. Since Christiaan Huygens first noted this phenomenon in the 17th century, various types of synchronization among weakly coupled oscillators, including firing neurons and flashing fireflies, have been documented by numerous physicists and biologists [3, 12, 24]. Systematic studies using mathematical models were later initiated by Arthur Winfree [30] and Yoshiki Kuramoto [23]. Among them, the phase synchronization model proposed by Kuramoto is considered as one of the prototype models for synchronization due to its simple mathematical form, the emergence of various synchronization phenomena, and its applicability to real-world cases such as Josephson junction arrays. The model proposed in [23], which is now called the Kuramoto model, is given as follows. Let θi be the phase of the i-th oscillator in R. The dynamics of the N phase configuration Θ:=(θ1,,θN) is governed by the system of ordinary differential equations:

θ˙i=νi+κNj=1Nsin(θj-θi),t>0, θ0i=θini,i[N]:={1,2,,N},

where the positive constant κ denotes the universal scale parameter of coupling strengths between each pair of particles, and νiR represents the natural frequency of the i-th oscillator in the absence of interaction, respectively.

As mentioned above, the Kuramoto model and its variations have been extensively studied in the literature as a prototype model for synchronization, particularly in relation to phase-transition behavior. For example, [1, 26] offer several numerical examples of the Kuramoto model from a physicist’s perspective. In [6, 9, 7, 8, 16, 22], the authors also studied variations of the Kuramoto model with additional structures, such as inertia or frustration. In [2, 18, 19], authors explored asymptotic phase and frequency synchronization for various initial conditions. Furthermore, [5, 10, 29, 28, 27] estimated the critical coupling strength needed to ensure the existence of phase-locked states under several conditions, including all-to-all and bipartite graph network topologies. Additionally, [13] provided sufficient conditions for the Kuramoto model with static network topology to exhibit asymptotic phase locking.

However, real-world systems exhibit inherent randomness and uncertainty that cannot be captured by deterministic models. Therefore, there have been many attempts to consider stochastically perturbed Kuramoto models to incorporate this variability, providing a more realistic representation. As far as we know, only stochastically perturbed Kuramoto models of the form

dθti=νi+κNj=1Nsin(θtj-θti)dt+j=1Nηij(t,Θt)dBtj,t>0, θ0i=θini,i[N],

have been considered so far, where the functions B1,,BN are independent and identically distributed (i.i.d.) standard one-dimensional Brownian motions, and ηij are bounded and Lipschitz continuous. For example, [11, 15] considered the additive white noise (ηij(t,Θt)=constant) to analyze its effect on the synchronization behavior of the Kuramoto model, and [17] used a multiplicative noise

ηij(t,Θt)=2σNk=1Nsin(θtk-θti)δ1j,i,j[N],

to make the equilibria of (3) asymptotically stable. In [20, 21] sufficient framework was provided to obtain

Plimt|θti-θtj|=0(modπ),i,j[N]=1,

when the phase configuration Θ satisfies

dθti=κNj=1Nsin(θtj-θti)dt+2σNj=1Nsin(θtj-θti)dBti,t>0, θ0i=θini,i[N],

which is the model that was originally intended to be analyzed in [17] instead of (4). The main difference between additive and multiplicative noise models lies in their behavior around equilibrium points. In additive noise models, the phase configuration generally stays near an equilibrium point for a long time, but eventually moves to other equilibrium point. In contrast, the equilibrium points that were stable in the deterministic setting are also asymptotically stable in multiplicative noise models. This difference arises because, in multiplicative noise models, the diffusion coefficient near the equilibrium point is close to zero, whereas it is not in additive noise models. Among all these results, there are no conditions that are necessary and sufficient for (1) or (4) to produce specific types of synchronization phenomena. Therefore, it is still meaningful to properly discretize the Kuramoto model, as numerical experiments with nonlinear differential equations often reveal facts about their solutions that cannot be proven due to technical limitations.

In this paper, we are interested in time discretization that preserves the asymptotic behavior of (4). More precisely, our interest lies in whether the already known asymptotic behavior of (4) can be accurately reproduced using the Euler-Maruyama method, one of the simplest methods for discretizing stochastic differential equations. Due to the similarity between drift and diffusion coefficients of (4), the approximated Euler-Maruyama solution for (4) can be obtained iteratively by using the recurrence relations

θn+1i=θni+κh+2σhZniNj=1Nsin(θnj-θni),i[N],nN{0}, θ0i=θini,Znii.i.d.N(0,1),i[N],nN{0}

where h:=Δt>0 is a fixed time step, and each Θn=(θn1,,θnN) approximates the solution of (4) at time nh. In addition, each Zni is a standard normal random variable where Zni and Zmj are independent if and only if nm or ij. For the model (4), we will consider an order parameter

R(Θ)=1Nj=1Neiθj

as an indicator of the phase synchronization, as it takes value in [0,1] and becomes 1 only if θ1==θN modulo 2π. Then, if there is a sufficient condition on κ,σ,h to satisfy

0E[R(Θ1)2]E[R(Θn)2]1,

we can say that (4) is synchronized in the sense that E[R2] increases monotonically. We will find such a condition by determining a function F=F(Θ) satisfying

E[R(Θn+1)2|Θn]-R(Θn)2F(Θn)0,

which make the series nE[F(Θn)] converges to a finite number. Then, by using the usual Borel-Cantelli argument, we will obtain

limnF(Θn)=0almost surely,

which indicates a specific synchronization phenomenon that the phase configuration will asymptotically exhibit a special shape. Although the use of the order parameter was also an important idea in the results of [17, 21, 20], the difference is that the Itô calculus cannot be used for the discrete-time model (4). Because of the nonlinear structure of R2 and the presence of unbounded noises Zni, it is impossible to approximate R(Θn+1)2-R(Θn)2 with the desired accuracy with probability 1 by Taylor approximation. Therefore, the main novelty of this paper is to show that even though we cannot control the individual sample paths of R(Θn)2, we can compute its expectation accurately and achieve the desired result.

The rest of this paper is organized as follows. In Section 2, we review several concepts and properties on (1) and (4), such as the order parameter and phase synchronization. In Section 3, we present a sufficient framework leading to the monotonic increasing property of E[R2] for (5). In Section 4, we verify our analysis through numerical experiments and discuss the optimality of the results. Finally, Section 5 is devoted to a brief summary of our results.

In this section, we present several well-known results on the deterministic Kuramoto model (1) and the stochastic Kuramoto model (4). First, we introduce the definition and some properties of order parameters, which are frequently used throughout our discussion.

Definition 2.1. Let Θ=(θ1,,θN) be a phase configuration in RN. Then, the order parameters (R(Θ),ϕ(Θ))R×R/2πZ are defined as

R(Θ):=1N j=1 Neiθj,eiϕ(Θ):=1NR(Θ) j=1Neiθj(if R(Θ)0). 

Since the above order parameters are well defined by the phase configuration alone, they can be thought of in a general context, not just for the Kuramoto model. To be more precise, in the case of φ, it can be said that the values of cosϕ and sinϕ are determined rather than just φ. Therefore, the following lemma provides a useful way to incorporate information about φ.

Lemma 2.2. [25] Let Θ=(θ1,,θN) be a phase configuration in RN. If R(Θ)>0, the order parameters (R(Θ),ϕ(Θ)) satisfy

R(Θ)cos(ϕ(Θ)-θ)=1Nk=1Ncos(θk-θ),R(Θ)sin(ϕ(Θ)-θ)=1Nk=1Nsin(θk-θ),

for every θR. In particular, we have

R(Θ)2=1Nj=1Ncos(θj-θ)2+1Nj=1Nsin(θj-θ)2θR, R(Θ)2=1N2i,j=1Ncos(θi-θj), R(Θ)=1Nj=1Ncos(θj-ϕ(Θ)),0=1Nj=1Nsin(θj-ϕ(Θ)).

Note that if R(Θ)=0, we can use k=1Neiθk=0 to obtain

1Nk=1Ncos(θk-θ)=Re1Nk=1Nei(θk-θ)=Re1Nk=1Neiθk·e-iθ=0,1Nk=1Nsin(θk-θ)=Im1Nk=1Nei(θk-θ)=Im1Nk=1Neiθk·e-iθ=0.

Therefore, although ϕ(Θ) is not well-defined in this case, (2) and (3) holds by replacing the undetermined number ϕ(Θ) to arbitrary constant.

Now, given the dynamics of the phase configuration, the derivative of R(Θ) can be expressed in terms of Θ by differentiating (3) and combining the derivative of Θ with (2). If we do this for the Kuramoto model, we obtain the following equation.

Proposition 2.3. Let Θ(t)=(θ1(t),,θN(t)) be a solution to the Kuramoto model (1). Then, R=R(Θ) satisfy

R˙=-1Nj=1Nνjsin(θj-ϕ)+κRNj=1Nsin2(θj-ϕ).

This is, in fact, a crucial estimate for the convergence of the Kuramoto model when all natural frequencies are identical. If νiν, we can use Lemma 2.2 and Proposition 2.3 to simplify (4) as

dR2dt=2κR2Nj=1Nsin2(θj-ϕ),

so that R is monotonically increasing. Therefore, if the initial configuration Θin satisfies R(Θin)>0, the phase ϕ(Θ)R/2πZ is well-defined. Then, since R is bounded above by 1, each sin(θj-ϕ) approaches zero as t tends to infinity, and we have the following result.

Proposition 2.4. [4] Let Θ(t)=(θ1(t),,θN(t)) be a solution to (1), where the natural frequencies {νi}i=1N and initial configuration Θin=(θin1,,θinN) satisfy

νiν,R(Θin)>0.

Then, the limit

θi:=limt(θi(t)-νt)

exists for all i[N] and

sin(θi-θj)=0,i,j[N].

When we consider the stochastic Kuramoto model (4), we assume there exists a probability triple (Ω,F,P), and denote (Ft)t0 the natural filtration generated by {Bt1,,BtN}. That is, Ft is the sub σ-algebra of F generated by {Bs1,,BsN:st}. Similar to Proposition 2.3, we can apply Itô's lemma to the stochastic Kuramoto model (4) as follows.

Proposition 2.5. [20, 21] Let Θt=(θt1,,θtN) be a solution process to the stochastic Kuramoto model (4). Then, Rt=R(Θt) satisfy

Ru2-Rs2=su[2κRt2Pt+2σRt2PtN-2σRt31Nj=1Nsin2(ϕt-θtj)cos(ϕt-θtj)]dt +22σNj=1NsuRt2sin2(ϕt-θtj)dBtj,

where the process Pt is given by

Pt:=1Ni=1Nsin2(θti-ϕt).

Therefore, by taking conditional expectation to both hand sides of (5), one can show that R(Θ)2 is a submartingale relative to ({Ft},P), provided that κ>1-1Nσ:

E[Ru2|Fs]=Rs2+suE[Rt2(2κPt+2σPtN-2σRt1Nj=1Nsin2(ϕt-θtj)cos(ϕt-θtj))]dt Rs2+suE[Rt2(2κPt+2σPtN-2σPt)]dt =Rs2+suE[Rt2Pt(2κ-2σ(1-1N))]dt Rs2,0su<.

In Section 3, we will provide a direct proof of the monotone increasing property of the expectation of R(Θn)2, without reproducing Proposition 2.5 in the discrete model (5). This result will lead to the conclusion that PlimnR(Θn)2P(Θn)=0=1.

In this section, we provide a synchronization estimate of (5). By using equation (2), we can rewrite the right hand side of (5) to

θn+1i=θni-(κh+2σhZni)Rnsin(θni-ϕn),i[N],nN{0}, θ0i=θini,Znii.i.d.N(0,1),i[N],nN{0}, Rn:=R(Θn),ϕn:=ϕ(Θn),

where we use the following relation instead of (1) when Rn=0:

Rn=0Rnsin(θni-ϕn)=0i[N].

Then, the idea is calculating the conditional expectation of R(Θn+1)2 for a fixed Θn. Once Θn is given, (1) shows that each θn+1i is a normal random variable, or just equal to θni if Rnsin(θni-ϕn). If we apply (3)2 in Lemma 2.2, Rn+12 can be considered as an average of resultant probability distributions obtained by taking the cosine of normal random variables. Therefore, we can use the following lemma to get E[Rn+12|Θn] explicity.

Lemma 3.1. For every λ,μR and a standard normal random variable Z, we have

E[cos(λZ+μ)]=e-λ22cosμ.

Proof. We use the fact that the characteristic function of the Gassian distribution is given by

E[eiλZ]=e-λ22

(see [14], p.210). Therefore, we have

E[cos(λZ+μ)]=E[Re(eiλZ+iμ)] =ReeiμE[eiλZ] =Ree-λ22(cosμ+isinμ) =e-λ22cosμ.

Now, using Lemma 3.1, (3)2 and introducing a simple notation, we can compute the conditional expectation as follows.

Lemma 3.2. Let (Θn)n0=(θn1,,θnN)n0 be a solution process to (1). Then, we have

E[Rn+12|Θn] =1N2N+jke-σhRn2sin2θ^nj+sin2θ^nkcosθ^nj-θ^nk-κhRn(sinθ^nj-sinθ^nk),

where the augmented phase θ^ni is defined as

θ^ni:=θni-ϕn,i[N],nN{0}.

Proof. First, let's fix j,k[N] and apply Lemma 3.1 to compute E[cos(θn+1j-θn+1k)|Θn]. From (1), we can represent θn+1j-θn+1k with respect to Θn and (Zn1,,ZnN) as follows:

θn+1j-θn+1k=θnj-(κh+2σhZnj)Rnsin(θnj-ϕn) -θnk-(κh+2σhZnk)Rnsin(θnk-ϕn) =θnj-θnk-κhRn(sin(θnj-ϕn)-sin(θnk-ϕn)) +2σhRn-sin(θnj-ϕn)Znj+sin(θnk-ϕn)Znk.

Since θnj-θnk=(θnj-ϕn)-(θnj-ϕk), we may use the augmented phases (2) to rewrite (3) as

θn+1j-θn+1k=θ^nj-θ^nk-κhRn(sinθ^nj-sinθ^nk) +2σhRn-sinθ^njZnj+sinθ^nkZnk.

If jk, the latter term in the right hand side of (4) is also a normal random variable with variance

(λnjk)2:=2σhRn2sin2θ^nj+sin2θ^nk,

as Znj and Znk are independent. Therefore, we denote

μnjk:=θ^nj-θ^nk-κhRn(sinθ^nj-sinθ^nk)

and apply Lemma 3.1 to obtain

E[cos(θn+1j-θn+1k)|Θn]=e-(λnjk)22cosμnjk =e-σhRn2sin2θ^nj+sin2θ^nkcosθ^nj-θ^nk-κhRn(sinθ^nj-sinθ^nk),

for every jk. Finally, using (3)2 in Lemma 2.2 yields

E[Rn+12|Θn] =E[1N2j,kcos(θn+1j-θn+1k)|Θn] =1N2(N+jkE[cos(θn+1j-θn+1k)|Θn]) =1N2N+jke-σhRn2sin2θ^nj+sin2θ^nkcosθ^nj-θ^nk-κhRn(sinθ^nj-sinθ^nk).

Next, we need to find the condition that makes E[Rn+12|Θn]Rn2. Since κ,σ, and N are constants that do not change and are variables used in the continuous-time model, we find a condition for the newly added variable h in the discrete-time model (1). In addition, since Θn varies depending on n, the desired condition should not depend on Θn. If we consider the explicit formula for E[Rn+12|Θn] in Lemma 3.2 as a function fn=fn(h) that depends only on h, we have

fn(0)=1N2N+jkcos(θ^nj-θ^nk) =1N21j,kNcos(θnj-θnk) =Rn2.

Then, we want to find the condition of h such that fn(h)fn(0) for every n, and for this we can use Taylor's theorem

fn(h)=fn(0)+hfn(0)+h22fn(h*),h*[0,h].

Therefore, in the following lemma, we find a positive constant h0 satisfying

  • h0=h0(κ,σ,N),

  • independent of n and Θn,

  • h0minh*[0,h0]2fn(0)|fn(h*)| for all n,

so that the relation E[Rn+12|Θn]Rn2 holds for every hh0.

Lemma 3.3. Let (Θn)n0=(θn1,,θnN)n0 be a solution process to (1), where κ,σ,N and h satisfy

κ>1-1Nσ,0<h2κ-21-1Nσκ2+21-1Nσ2.

Then, we have the following inequality: for every nN{0},

E[Rn+12|Θn]Rn2+h2κ-21-1Nσ-hκ2+21-1Nσ2Rn2Pn Rn2.

Proof. From now, we use the following notation for simplicity:

cosθ^ni=c^ni,sinθ^ni=s^ni,i[N],nN{0}.

Then, we can write fn(h) as

fn(h)=1N2N+ije-σhRn2(s^ni)2+(s^nj)2cosθ^ni-θ^nj-κhRn(s^ni-s^nj).

By differentiating fn with respect to h, one can obtain

fn(h)=1N2ije-σhRn2((s^ni)2+(s^nj)2)   [-σRn2((s^ni)2+(s^nj)2)cos(θ^ni-θ^nj-κhRn(s^ni-s^nj)) +κRn(s^ni-s^nj)sin(θ^ni-θ^nj-κhRn(s^ni-s^nj))] =1N2ije-σhRn2((s^ni)2+(s^nj)2)gnij(h),

where gnij(h) is defined as

gnij(h):=-σRn2((s^ni)2+(s^nj)2)cos(θ^ni-θ^nj-κhRn(s^ni-s^nj)) +κRn(s^ni-s^nj)sin(θ^ni-θ^nj-κhRn(s^ni-s^nj)).

Now, we are ready to estimate fn(0) and fn(h).

(1) For fn(0), the direct computation yields

fn(0)=1N2ijgnij(0) =1N2ij[-σRn2((s^ni)2+(s^nj)2)cos(θ^ni-θ^nj)+κRn(s^ni-s^nj)sin(θ^ni-θ^nj)] =1N2ij[-σRn2((s^ni)2+(s^nj)2)(c^nic^nj+s^nis^nj)+κRn(s^ni-s^nj)(s^nic^nj-c^nis^nj)] =1N21i,jN[-σRn2((s^ni)2+(s^nj)2)(c^nic^nj+s^nis^nj)+κRn(s^ni-s^nj)(s^nic^nj-c^nis^nj)] +2σRn2N2i=1N(s^ni)2.

Then, we use (3)3 of Lemma 2.2, i.e.,

1Ni=1Nc^ni=Rn,1Ni=1Ns^ni=0,

to obtain

fn(0)=1N21i,jN[-σRn2((s^ni)2+(s^nj)2)(c^nic^nj)+κRn((s^ni)2c^nj+(s^nj)2c^ni)] +2σRn2N2i=1N(s^ni)2 =2κRn2Ni=1N(s^ni)2-2σRn3Ni=1N(s^ni)2c^ni+2σRn2N2i=1N(s^ni)2.

This exactly coincides with the drift part of (4), confirming that all calculations up to this point are correct. Therefore, if we use the notation

Pn:=1Ni=1N(s^ni)2,

we have

fn(0)2Rn2Pnκ-σRn+1Nσ.

(2) For fn(h), we first represent fn(h) in terms of gnij and (gnij) as follows:

fn(h)=1N2ije-σhRn2((s^ni)2+(s^nj)2)[gnij(h)·(-σRn2((s^ni)2+(s^nj)2))+(gnij)(h)],

where each (gnij)(h) is given by

(gnij)(h)=-κσRn3((s^ni)2+(s^nj)2)(s^ni-s^nj)sin(θ^ni-θ^nj-κhRn(s^ni-s^nj)) -κ2Rn2(s^ni-s^nj)2cos(θ^ni-θ^nj-κhRn(s^ni-s^nj)).

To compute an upper bound of (8), we again introduce the following notation:

Cnij(h):=cos(θ^ni-θ^nj-κhRn(s^ni-s^nj)),Snij(h):=sin(θ^ni-θ^nj-κhRn(s^ni-s^nj)).

Then, we combine (4) and (9) to obtain

gnij(h)·(-σRn2((s^ni)2+(s^nj)2))+(gnij)(h) =(-σRn2((s^ni)2+(s^nj)2)Cnij(h)+κRn(s^ni-s^nj)Snij(h))·(-σRn2((s^ni)2+(s^nj)2)) -κσRn3((s^ni)2+(s^nj)2)(s^ni-s^nj)Snij(h)-κ2Rn2(s^ni-s^nj)2Cnij(h) =σ2Rn4((s^ni)2+(s^nj)2)2-κ2Rn2(s^ni-s^nj)2Cnij(h) -[2κσRn3((s^ni)2+(s^nj)2)(s^ni-s^nj)]Snij(h).

Since |ACnij(h)+BSnij(h)|A2+B2 for every A,BR, we have

|gnij(h)·(-σRn2((s^ni)2+(s^nj)2))+(gnij)(h)| σ2Rn4((s^ni)2+(s^nj)2)2-κ2Rn2(s^ni-s^nj)22+[2κσRn3((s^ni)2+(s^nj)2)(s^ni-s^nj)]2 =σ2Rn4((s^ni)2+(s^nj)2)2+κ2Rn2(s^ni-s^nj)2.

Therefore, we apply this inequality to (8) to obtain

|fn(h)|1N2ij[gnij(h)·(-σRn2((s^ni)2+(s^nj)2))+(gnij)(h)] 1N2ij[σ2Rn4((s^ni)2+(s^nj)2)2+κ2Rn2(s^ni-s^nj)2] =1N21i,jN[σ2Rn4((s^ni)2+(s^nj)2)2+κ2Rn2(s^ni-s^nj)2]-1N2i=1N4σ2Rn4(s^ni)4 =σ2Rn41N21i,jN(s^ni)4+2(s^ni)2(s^nj)2+(s^nj)4-4N2i=1N(s^ni)4 +κ2Rn21N21i,jN(s^ni)2-2s^nis^nj+(s^nj)2 =σ2Rn42-4N1Ni=1N(s^ni)4+2Pn2+κ2Rn22Pn,

where we used

1Ni=1Ns^ni=0,

in the last equality.Then, we combine (7) and (10) to obtain

2fn(0)|fn(h*)|4Rn2Pnκ-σRn+1Nσσ2Rn42-4N1Ni=1N(s^ni)4+2Pn2+2κ2Rn2Pn 4Rn2Pnκ-σRn+1Nσσ2Rn42-4NPn+2Pn2+2κ2Rn2Pn =2κ-σRn+1Nσκ2+σ2Rn21-2N+Pn 2κ-21-1Nσκ2+21-1Nσ2,

for every h*R. Finally, we apply (7) and (11) to obtain

E[Rn+12|Θn]=fn(h) =fn(0)+hfn(0)+h22fn(h*) fn(0)+h1-κ2+21-1Nσ22κ-21-1Nσ·hfn(0) fn(0)+h2κ-21-1Nσ-hκ2+21-1Nσ2Rn2Pn.

for all nN{0}, which is our desired result.

Remark 3.4. In [25], the author proved that the discrete deterministic Kuramoto model

θn+1i=θni-κhRnsin(θni-ϕn),i[N],nN{0}, θ0i=θini,i[N],

has a monotone increasing property for Rn2 when κ>0 and 0<κh2. Thus, the result of Lemma 3.3 is consistent with [25] as σ tends to 0.

Consequently, we can show the convergence of the series nE[Rn2Pn]. This not only implies that Rn2Pn will converge to zero almost surely, but also leads to all of the following results.

Theorem 3.5. Let (Θn)n0=(θn1,,θnN)n0 be a solution process to (1), where κ,σ,N and h satisfy

κ>1-1Nσ,0<h<2κ-21-1Nσκ2+21-1Nσ2.

Then, we have

PlimnRns^ni=0=PlimnRns^niZnj=0=Plimn|θn+1i-θni|=0=1,i,j[N].

In addition, there exists a random variable R:Ω[0,1] satisfying

PlimnRn=R=1.

Proof. First, we set a constant independent of n:

C(κ,σ,N,h):=h2κ-21-1Nσ-hκ2+21-1Nσ2>0.

Then, we apply the tower property of the conditional expectation to (4) to obtain

E[Rn+12]=EE[Rn+12|Θn] ERn2+C(κ,σ,N,h)Rn2Pn =E[Rn2]+C(κ,σ,N,h)E[Rn2Pn],nN{0}.

Since the order parameter Rn is always containted in [0,1], this yields

C(κ,σ,N,h)n=0E[Rn2Pn]n=0(E[Rn+12]-E[Rn2])1-Rin2<.

Note that Θn and Zni are independent for every i[N] and nN{0}. Thus, we can also obtain

n=0E[Rn2Pn]=n=0E[Rn2Pn(Zni)2]<,i[N].

Then, we use the Markov inequality to get

n=1PRn2Pnε1εn=1E[Rn2Pn]<,ε>0, n=1PRn2Pn(Zni)2ε1εn=1E[Rn2Pn(Zni)2]<,ε>0.

By using the Borel-Cantelli lemma, (12) implies that

Pn=1k=n{Rk2Pkε}=0,Pn=1k=n{Rk2Pk(Zki)2ε}=0,i[N],ε>0,

which shows the P-almost sure convergence of Rk2Pk and Rk2Pk(Zki)2 to 0. Since Rn2Pn and Rn2Pn(Znj)2 can be written as

Rn2Pn=1Ni=1N(Rns^ni)2,Rn2Pn(Znj)2=1Ni=1N(Rns^niZnj)2,

we have the desired P-almost sure convergence of Rns^ni and Rns^niZnj as n. In addtion, we apply these almost sure convergence results to (1) to obtain

Plimn|θn+1i-θni|=0=PlimnRns^ni|κh+2σhZni|=0=1.

Finally, since Rn2 is a bounded submartingale (relative to natural filtration generated by Θn), we apply Doob's supermartingale convergence theorem to show the existence of finite random variable R:=limnRn.

Remark 3.6. Suppose the system parameters κ,σ,N,h satisfy the assumption of Theorem 3.5. If R(Θin)>0, the expectation of R2 must be positive due to the monotone increasing property of E[Rn2]. Therefore, there are positive measures of ωΩ such that

limnRn(ω)=R(ω)>0,limnR(Θn(ω))s^ni(ω)=0,i[N].

If all s^ni(ω) converges to zero as n tends to infinity, we have

limnsin(θni(ω)-ϕn(ω))=0,i[N],

and this leads to the following conclusion: for every i,j[N], we have

limnsin(θni-θnj)=0.

This result corresponds to the synchronization of deterministic Kuramoto model, for instance the result in Proposition 2.4.

In this section, we perform numerical simulations for the stochastic Kuramoto model, particularly using the Euler-Maruyama method (4). All simulations were conducted over the time interval t[0,20], with the system parameters N and κ fixed as follows:

N=10,h=0.01,κ=1.

We also fixed Θin by selecting each initial θini randomly from the uniform distribution U[0,2π]. Thus, we only compare how the dynamics of (4) vary with respect to the noise strength σ. To estimate E[Rn2], we used the empirical mean of m=300 simulations for stochastic processes Rn2 and denote R¯n2. Then, the following Figure 1 shows the temporal evolutions of R¯n2 for

Figure 1. R¯n2 vs nh.

σ=0,1,2,4,10.

Since κh=0.012, the R2 of the case σ=0 increases monotonically (see Remark 3.4) and approaches to 1 as t, which means that every θi-θj indeed converged to 0 modulo 2π. The case σ=1 also satisfies the condition presented in Theorem 3.5, so that R¯n2 increases monotonically as theoretically expected. However, all other R¯n2 where σ does not satisfy the condition of Theorem 3.5 also seem to exhibit the monotone increasing property. To analyze why this happened, let's recall the linear approximation of fn(h)=E[Rn+12|Θn] obtained from Lemma 3.3, which is as follows:

fn(0)=2κRn2Ni=1N(s^ni)2-2σRn3Ni=1N(s^ni)2c^ni+2σRn2N2i=1N(s^ni)2.

In Lemma 3.3, we used the inequality

Rn3Ni=1N(s^ni)2c^niRn2Ni=1N(s^ni)2,

and then a condition κ>1-1Nσ ws required to ensure that fn(0) is strictly positive. The most likely reason the condition appears to be non-optimal is that (1) is suboptimal in many cases. First, if the order parameter R is smaller than a constant r<1, the left-hand side of (1) must be at most r times the right-hand side. Secondly, if all c^ni's are positive, the order of magnitude of (s^ni)2 and c^ni are the opposite. From the perspective of the rearrangement inequality, the average of (s^ni)2c^ni is smaller than the product of the average of (s^ni)2 (which is Pn) and the average of c^ni (which is Rn).

Next, we discuss each individual sample path of Rn2, instead of its expectation E[Rn2]. In Figure 2, the values of Rn2 are plotted over n by performing simulations 10 times each when σ=1,2,4 and σ=10. Here, we see that the value of Rn2 remains almost constant after reaching certain values. There are a total of six ways to divide the 10 phases into two groups: 10:0, 9:1, 8:2, 7:3, 6:4, and 5:5. If k{0,1,2,3,4,5} and

Figure 2. Sample paths of Rn2

θ1==θk=θk+1+π==θ10+πmodulo2π,

the order parameter R becomes (10-k)-k10=10-2k10{0,0.2,0.4,0.6,0.8,1}, which we call the bipolar state. As a result, R2 remains almost constant in the vicinity of R2=0.16,0.36,0.64 and 1 in Figure 2. However, in all cases except when σ=1, they are not completely trapped in such regions due to the influence of random noise, and there exists a sample path going from one bipolar state to another. In the transition between bipolar states, the probability of moving toward a larger R2 seems generally higher. This could be another reason for the increasing property of E[Rn2] in Figure 1 when σ is larger than the upper bound presented in Theorem 3.5.

In this paper, we have introduced a direct approach to examine the emergence of synchronization in the discrete-time stochastic Kuramoto model. The sufficient conditions proposed in Theorem 3.5 are consistent with both the results for the continuous stochastic model [17, 21] and the discrete deterministic Kuramoto model [25]. However, the numerical simulations in Section 4 suggest that the sufficient conditions presented in [21] may be too strong. Based on the findings of this paper and existing literature, we believe the following research topics can be studied in future:

  • Stability of bipolar states for large values of σκ.

  • Transition probabilities from one bipolar state to another.

  1. J. A. Acebrón, L. L. Bonilla, C. J. Pérez Vicente, F. Ritort and R. Spigler. The Kuramoto model: A simple paradigm for synchronization phenomena, Rev. Mod. Phys, 77(2005), 137-185. ArXiv: 2005.9547v2.
    CrossRef
  2. D. Aeyels and J. Rogge. Stability of phase locking and existence of frequency in networks of globally coupled oscillators, Prog. Theor. Phys., 112(2004), 921-941.
    CrossRef
  3. J. Buck and E. Buck. Biology of synchronous flashing of fireflies, Nature, 211(1966), 562-564.
    CrossRef
  4. D. Benedetto, E. Caglioti and U. Montemagno. On the complete phase synchronization for the Kuramoto model in the mean-field limit, Commun. Math. Sci, 13(2015), 1775-1786.
    CrossRef
  5. J. C. Bronski, L. Deville and M. J. Park. Fully synchronous solutions and the synchronization phase transition for the finite-N Kuramoto model, Chaos, 22(2012), 033133.
    Pubmed CrossRef
  6. Y.-P. Choi, S.-Y. Ha and J. Morales. Emergent dynamics of the Kuramoto ensemble under the effect of inertia, Discrete Contin. Dyn. Syst., 38(2018), 4875-4913.
    CrossRef
  7. Y.-P. Choi, S.-Y. Ha and S. Noh. Remarks on the nonlinear stability of the Kuramoto model with inertia, Quart. Appl. Math, 73(2015), 391-399.
    CrossRef
  8. Y.-P. Choi, S.-Y. Ha and S.-B. Yun. Global existence and asymptotic behavior of measure valued solutions to the kinetic Kuramoto-Daido model with inertia, Netw. Heterog. Media, 8(2013), 943-968.
    CrossRef
  9. Y.-P. Choi, Z. Li, S.-Y. Ha, X. Xue and S.-B. Yun. Complete entrainment of Kuramoto oscillators with inertia on networks via gradient-like flow, J. Differential Equations, 257(2014), 2591-2621.
    CrossRef
  10. N. Chopra and M. W. Spong. On exponential synchronization of Kuramoto oscillators, IEEE Trans. Automat. Control, 54(2009), 353-357.
    CrossRef
  11. L. DeVille. Transitions amongst synchronous solutions in the stochastic Kuramoto model, Nonlinearity, 25(2012), 1473-1494.
    CrossRef
  12. F. Dörfler and F. Bullo. Synchronization in complex networks of phase oscillators: A survey, Automatica J. IFAC, 50(2014), 1539-1564.
    CrossRef
  13. J.-G. Dong and X. Xue. Synchronization analysis of Kuramoto oscillators, Commun. Math. Sci, 11(2013), 465-480.
    CrossRef
  14. B. Fristedt and L. Gray, A Modern Approach to Probability Theory. Probability and Its Applications, Birkhäuser, Boston, (1997).
    Pubmed CrossRef
  15. B. Gentz, S.-Y. Ha, D. Ko and C. Wiesel. Emergent dynamics of Kuramoto oscillators under the effect of additive white noises, Preprint.
  16. S.-Y. Ha, D. Kim, J. Lee and Y. Zhang. Remarks on the stability properties of the Kuramoto-Sakaguchi-Fokker-Planck equation with frustration, Z. Angew. Math. Phys, 69(2018), 25 pp.
    CrossRef
  17. S.-Y. Ha, D. Ko, C. Min and X. Zhang. Emergent collective behaviors of stochastic Kuramoto oscillators, Discrete Contin. Dyn. Syst. Ser. B, 25(2020), no.3, 1059-1081.
  18. S.-Y. Ha, H. K. Kim and S. Ryoo. Emergence of phase-locked states for the Kuramoto model in a large coupling regime, Commun. Math. Sci, 14(2016), 1073-1091.
    CrossRef
  19. S.-Y. Ha, H. K. Kim and S. Ryoo. On the Finiteness of Collisions and phase-locked States for the Kuramoto Model, J. Stat. Phys, 163(2016), 1394-1424.
    CrossRef
  20. S.-Y. Ha, D. Ko and W. Shim. On the stochastic synchronization of the Winfree model with a multiplicative noise, Commun. Math. Sci., 22(2024), 1955-1983.
    CrossRef
  21. S.-Y. Ha, D. Ko, W. Shim and H. Yu. Emergent behaviors of the Justh-Krishnaprasad model with uncertain communications, J. Differential Equations, 322(2022), 38-70.
    CrossRef
  22. S.-Y. Ha and Z. Li. Uniqueness and well-ordering of emergent phase-locked states for the Kuramoto model with frustration and inertia, Math. Models Methods Appl. Sci, 26(2016), 357-382.
    CrossRef
  23. Y. Kuramoto. Self-entrainment of a population of coupled non-linear oscillators, International Symposium on Mathematical Problems in Mathematical Physics, Lecture Notes in Phys., 30(1975), pp. 420-422..
    CrossRef
  24. A. Pikovsky, M. Rosenblum and J. Kurths, Synchronization: A universal concept in nonlinear sciences, Cambridge University Press, Cambridge, (2001).
    CrossRef
  25. W. Shim. On the generic complete synchronization of the discrete Kuramoto model, Kinet. Relat. Models, 13(2020), 979-1005.
    CrossRef
  26. S. H. Strogatz. From Kuramoto to Crawford: exploring the onset of synchronization in populations of coupled oscillators, Phys. D, 143(2000), 1-20.
    CrossRef
  27. M. Verwoerd and O. Mason. Global phase-locking in finite populations of phase-coupled oscillators, SIAM J. Appl. Dyn. Syst, 7(2008), 134-160.
    CrossRef
  28. M. Verwoerd and O. Mason. On computing the critical coupling coefficient for the Kuramoto model on a complete bipartite graph, SIAM J. Appl. Dyn. Syst, 8(2009), 417-453.
    CrossRef
  29. M. Verwoerd and O. Mason. A convergence result for the Kuramoto model with all-to-all coupling, SIAM J. Appl. Dyn. Syst, 10(2011), 906-920.
    CrossRef
  30. A. T. Winfree. Biological rhythms and the behavior of populations of coupled oscillators, J. Theor. Biol., 16(1967), 15-42.
    Pubmed CrossRef